cond-mat0010165/ps4.tex
1: \documentstyle[prb,aps,psfig]{revtex}
2: %\documentstyle[preprint,aps,psfig]{revtex}
3: %\input epsf.sty
4: \begin{document}
5: \draft
6: %%%%%%%%%%%%%%%%%%
7: \twocolumn[\hsize\textwidth\columnwidth\hsize\csname
8: @twocolumnfalse\endcsname
9: %%%%%%%%%%%%%%%%%%
10: 
11: 
12: \title{Spatially homogeneous ground state of the two-dimensional Hubbard 
13: model}
14: \author{Federico Becca, Massimo Capone, and Sandro Sorella}
15: \address{
16: Istituto Nazionale per la Fisica della Materia and 
17: International School for Advanced Studies, Via Beirut 4,
18: 34013 Trieste, Italy\\}
19: 
20: \date{\today}
21: \maketitle
22: \begin{abstract}
23: We investigate the stability with respect to phase separation or charge
24: density-wave formation of
25: the two-dimensional Hubbard model for various values of the local Coulomb
26: repulsion and electron densities using Green-function Monte
27: Carlo techniques.
28: The well known sign problem is particularly serious in 
29: the relevant region of small hole doping. 
30: We show that the difference in accuracy for different doping makes
31: it very difficult to probe the phase separation instability using only
32: energy calculations, even in the weak-coupling limit ($U=4t$) 
33: where reliable results are available.
34: By contrast, the knowledge of the charge correlation functions allows us to
35: provide clear evidence of a spatially homogeneous ground state up to $U=10t$.
36: \end{abstract}
37: \pacs{71.10.Fd, 71.45.Lr, 74.20.-z}
38: %%%%%%%%%%%%%%%%%%
39: ]
40: %%%%%%%%%%%%%%%%%%
41: 
42: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
43: %                               TEXT
44: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45: \section{Introduction}
46: 
47: Since the discovery of high-temperature superconductivity, the two-dimensional
48: Hubbard model \cite{hubbard} has been the subject of a
49: huge amount of work. Indeed, it is widely accepted that, despite the 
50: complicate structure of these materials, a key role is played by the 
51: electron correlation in the ${\rm CuO}$ planes, well represented by 
52: the Hubbard model.
53: 
54: This model and the related $t-J$ model reproduce, at least
55: qualitatively, some of the physical properties of cuprates. For 
56: example, at half-filling (i.e., when the number of electrons is equal to the
57: number of sites) the ground state (GS) is an antiferromagnetic insulator,
58: and upon doping the antiferromagnetism is strongly suppressed. 
59: The possible insurgence of superconductivity upon doping and the 
60: symmetry of the order parameter in the model are still 
61: open questions.
62: Another important point to study is the stability with respect to 
63: phase separation (PS) or charge-density waves (CDW) near half-filling. 
64: Indeed, both PS and CDW have been experimentally observed in different 
65: systems \cite{jorgensen,tranquada} and different authors 
66: \cite{emery,castellani} have pointed out
67: a possible relation between charge instability and superconductivity.
68: 
69: Because of the strongly interacting nature of these systems, important
70: insights came from nonperturbative numerical methods.
71: 
72: The exact diagonalization methods, e.g., Lanczos, are strongly limited 
73: by the exponential 
74: growth of the Hilbert space, and in practice it is possible to diagonalize 
75: only up to about 20 sites. 
76: A remarkable development of the exact diagonalization strategy is the density
77: matrix renormalization group \cite{white}, which allows us to compute the
78: GS using an iteratively improved basis.
79: Despite the accuracy of this technique for one-dimensional
80: and quasi-one-dimensional systems, there is no straightforward 
81: generalization for higher dimensions.
82: 
83: Attractive alternatives are stochastic methods such as
84: Quantum Monte Carlo (QMC) that do not suffer from severe limitations
85: in terms of lattice size and hence allow us to study fairly large systems.
86: The simplest QMC simulation is the variational one, in which it is 
87: possible to compute the expectation values of the energy and 
88: all the correlation functions for a given wave function. 
89: The limitation of this method is that one has to guess {\it a priori}
90: the form of the GS.
91: 
92: QMC methods allow us to sample directly the GS by using projection
93: techniques, the simplest one being the so-called power method, namely
94: the simple idea that an iterative application of the Hamiltonian $H$
95: filters out the ground state starting from an arbitrary state
96: nonorthogonal to it. These approaches suffer, in fermionic systems,
97: from the so-called sign problem. Indeed, due to the
98: antisymmetry of the wave function under permutations of two particles,
99: one gets after a few power iterations opposite contributions leading
100: to the cancellation of large and fluctuating weights. 
101: 
102: An alternative approach, which is quite efficient for the Hubbard model,
103: is the auxiliary field quantum Monte Carlo (AFQMC) \cite{afqmc}.
104: The Coulomb interaction is linearized by using a Hubbard-Stratonovic
105: transformation. In this way it is possible to apply exactly the
106: operator $e^{-\beta H}$ to a Slater determinant.
107: In practice, AFQMC is very accurate only for small values
108: of the Coulomb repulsion, whereas, when the interaction 
109: strength becomes comparable with
110: the bare bandwidth, large fluctuations make the simulations rather
111: unstable and inefficient, because the sign problem becomes 
112: particularly severe for strong Coulomb repulsion.
113: 
114: A sophisticated implementation of the power method 
115: is the Green-function Monte Carlo (GFMC) applied to lattice models
116: \cite{nandini}. 
117: In principle, this method gives unbiased results but in
118: practice, for fermion systems, the sign problem makes 
119: any simulation prohibitive, because of statistical errors. 
120: Recently, many attempts have been made to deal with fermions
121: and overcome the sign problem \cite{cpqmc,fn}. 
122: The fixed-node approximation \cite{fn} 
123: is introduced for the  GFMC and replaces the original Hamiltonian
124: with an effective one free of the 
125: sign problem, in which the nodes are fixed to be those 
126: of the so-called guiding wave function.
127: In practice, the most common choice is the best variational
128: wave function available, i.e., the one with the lowest energy. 
129: The fixed-node energy is an upper bound to the true GS energy and 
130: the method is variational.
131: Moreover, it becomes exact if the guiding function is
132: chosen to be the exact GS function. Although this 
133: approximation can be uncontrolled and needs an external input, 
134: it is quite accurate with respect to both energy and correlation functions 
135: for small sizes. 
136: 
137: An alternative approach, 
138: recently proposed by one of us, is to use the GFMC with a stochastic
139: reconfiguration (SR) \cite{sorella}.
140: The main idea of this method is to use a suitable reference dynamics 
141: free of the sign problem to constrain the true dynamics.  
142: The fixed-node and a variational dynamics are natural 
143: candidates as references.
144: At each reconfiguration the amplitudes of the GS
145: are chosen as small perturbations of the reference wave function
146: amplitudes such that the mixed averages of a few physically
147: relevant quantities are conserved. The method is exact if
148: all the operators in the Hilbert space are reconfigured, but it has been
149: shown that a major role is played by a very small number of 
150: operators \cite{sorcap}.
151: 
152: In this paper,
153: we present a systematic study on the Hubbard model comparing different 
154: QMC techniques by studying in particular the stability of PS 
155: and CDW near half-filling
156: for various values of the Coulomb repulsion.
157: In Sec. \ref{model} we introduce the Hubbard model and we discuss
158: the implementation of our approximations, in Sec. 
159: \ref{results} we report our results,
160: and finally in Sec. \ref{conclusions} we give a brief summary.
161: 
162: \section{The Model}\label{model}
163: 
164: We consider the Hubbard model on a square lattice of $L$ sites with 
165: $N=N_{\uparrow}+N_{\downarrow}$ particles and  $N_{\uparrow}=N_{\downarrow}$,
166: where $N_{\uparrow}$ ($N_{\downarrow}$) is the number of spin-up (-down) particles.
167: In order to study PS as close as possible to half-filling 
168: using only closed-shell configurations, we consider
169: square lattices tilted by $45^\circ$ with $L^2=2 l^2$ and $l$ odd.
170: In this way half-filling is a closed shell and the first doped closed shell
171: has eight holes independent of $L$.
172: The Hamiltonian reads
173: \begin{equation}\label{hamilt}
174: H = -t \sum_{\langle i,j \rangle, \sigma} c^{\dag}_{i,\sigma} c_{j,\sigma}
175: + U \sum_{i} n_{i,\uparrow} n_{i,\downarrow},
176: \end{equation}
177: where $\langle \; \rangle$ stands for nearest neighbors, $c_{i,\sigma}$ 
178: ($c^{\dag}_{i,\sigma}$) destroys (creates) an electron with spin $\sigma$ at
179: site $i$, and $n_{i,\sigma}=c^{\dag}_{i,\sigma}c_{i,\sigma}$. 
180: In the  following, all energies are measured in units of $t$. 
181: 
182: As already pointed out above, in the presence of the sign problem 
183: the choice of the guiding wave function is crucial.
184: Our wave function reads
185: \begin{equation}\label{afwf}
186: |\Psi_{G} \rangle = {\cal P}_{Sz=0} {\cal P}_g {\cal J} |{\cal D} \rangle ,
187: \end{equation}
188: where $|{\cal D} \rangle$ is a Slater determinant in which the orbitals are suitably 
189: chosen (see below), ${\cal P}_{Sz=0}$ is the projector onto the
190: subspace with $N_{\uparrow}=N_{\downarrow}$, i.e., with zero total spin
191: component along the $z$ axis,
192: ${\cal P}_g$ is a Gutzwiller operator that inhibits 
193: the double occupancies ${\cal P}_g = exp(-g\sum_{i} n_{i\uparrow}n_{i\downarrow})$,
194: where $g$ is a variational parameter, and $\cal J$ is a Jastrow factor
195: ${\cal J} = exp(\frac{\gamma}{2}\sum_{i,j} v_{i,j} S_i^z S_j^z)$, 
196: where $\gamma$ is another variational parameter and $v_{i,j}$ is taken from 
197: spin-waves theory \cite{franjic}.
198: Care must be taken in the choice of the orbitals appearing in 
199: the Slater determinant.
200: The most common choice is to take
201: the orbitals from a Hartree-Fock (HF) approximation of the Hamiltonian
202: breaking the SU(2) spin rotation symmetry along the $z$ axis. 
203: \begin{equation}\label{hf}
204: H = -t \sum_{\langle i,j \rangle, \sigma} c^{\dag}_{i,\sigma} c_{j,\sigma}
205: + \frac{U}{2} \sum_{i, \sigma} 
206: \left [ 
207: \langle n_i \rangle - \sigma (-1)^{R_i} \langle m_i \rangle
208: \right ] n_{i,\sigma},
209: \end{equation}
210: where
211: \begin{eqnarray}
212: \langle n_i \rangle & = & \langle n_{i,\uparrow} \rangle +
213: \langle n_{i,\downarrow} \rangle, \\
214: \langle m_i \rangle & = & (-1)^{R_i} \left [
215: \langle n_{i,\uparrow} \rangle -
216: \langle n_{i,\downarrow} \rangle \right ].
217: \end{eqnarray}
218: We consider only fillings which are closed shells for $U=0$ 
219: and where a solution with constant density 
220: $\langle n_i \rangle=\frac{N}{L}=n$ and staggered magnetization 
221: $\langle m_i \rangle=m$ are found.
222: In this case, the HF many-body wave function can be written as
223: \begin{equation}\label{det1}
224: |{\cal D} \rangle = 
225: \prod^{1,...,N_{\uparrow}}_k \beta^{\dag}_{k,\uparrow}
226: \prod^{1,...,N_{\downarrow}}_q \beta^{\dag}_{q,\downarrow} |0 \rangle,
227: \end{equation}
228: where the quasiparticles have definite momentum modulo $Q=(\pi,\pi)$ and
229: definite spin, since 
230: the antiferromagnetic order parameter is along the $z$ axis
231: \begin{equation}
232: \beta^{\dag}_{k,\sigma}=
233: u_k c^{\dag}_{k,\sigma}+\sigma v_k c^{\dag}_{k+Q,\sigma},
234: \end{equation}
235: $k$ is in the reduced magnetic Brillouin zone, and
236: $u_k$ and $v_k$ are defined in Ref.~\cite{shiba}. It is worth noting that
237: for $U/t \rightarrow \infty$, $u_k=v_k=\frac{1}{\sqrt{2}}$, namely the
238: spin up and the spin down are in different sub-lattices 
239: (classical N\'eel state).
240: 
241: In a previous work \cite{cosentini} the wave function 
242: (\ref{afwf}) with $|{\cal D} \rangle$ given by Eq.~(\ref{det1}) has been 
243: found to be a rather poor approximation for large $U/t$ at half-filling.
244: In particular, in this representation the Jastrow factor ${\cal J}$ does
245: not play any important role.
246: 
247: We propose a wave function which is a straightforward generalization
248: of the one successfully used for the Heisenberg model \cite{calandra}.
249: The fundamental ingredient is to allow spin 
250: fluctuations perpendicular to the staggered magnetization. An easy 
251: implementation of this idea is to put the magnetization 
252: in the $x-y$ plane allowing transverse fluctuations along the $z$ axis 
253: through a Jastrow-like factor \cite{franjic}.
254: This is achieved by a $\frac{\pi}{2}$ rotation $U_y(\frac{\pi}{2})$ 
255: around the $y$ axis of
256: the canonical operators:
257: \begin{eqnarray}
258: U^{\dag}_y(\frac{\pi}{2}) c^{\dag}_{i,\uparrow} U_y(\frac{\pi}{2}) & = & 
259: \frac{1}{\sqrt{2}} 
260: \left ( c^{\dag}_{i,\uparrow} + c^{\dag}_{i,\downarrow} \right ), \\
261: U^{\dag}_y(\frac{\pi}{2}) c^{\dag}_{i,\downarrow} U_y(\frac{\pi}{2}) & = & 
262: \frac{1}{\sqrt{2}} 
263: \left ( c^{\dag}_{i,\uparrow} - c^{\dag}_{i,\downarrow} \right ).
264: \end{eqnarray}
265: The fermionic part of our guiding wave function is therefore defined 
266: as a Slater determinant of the transformed orbitals,
267: \begin{eqnarray}
268: \label{rotated}
269: \beta^{\dag}_{k,+} &=& 
270: U^{\dag}_y(\frac{\pi}{2}) \beta^{\dag}_{k,\uparrow} U_y(\frac{\pi}{2}), \\
271: \label{rotated2}
272: \beta^{\dag}_{k,-} &=&
273: U^{\dag}_y(\frac{\pi}{2}) \beta^{\dag}_{k,\downarrow} U_y(\frac{\pi}{2}),
274: \end{eqnarray}
275: namely it is given by
276: \begin{equation}\label{det2}
277: |{\cal D} \rangle = 
278: \prod^{1,...,N_{\uparrow}}_k \beta^{\dag}_{k,+}
279: \prod^{1,...,N_{\downarrow}}_q \beta^{\dag}_{q,-} |0 \rangle.
280: \end{equation}
281: Remarkably for $U/t \rightarrow \infty$ and half-filling, by construction 
282: Eq.~(\ref{det2}) becomes the N\'eel state with spin
283: quantization parallel to the $x$ axis, i.e. it has the correct
284: Marshall sign on each of the $2^L$  configurations sampled by GFMC.
285: In this limit, it is also clear why the Jastrow factor may be much more 
286: effective: being defined along the $z$ axis, 
287: it allows us to sample the quantum fluctuation perpendicular to 
288: the staggered magnetization. In the previous case instead both  the Jastrow 
289: quantization axis and the order parameter were 
290: parallel, and for $U/t \rightarrow \infty$ there is no way to sample any
291: fluctuation, the only possible configuration being  the classical one.
292: 
293: It is worth noting that, although for the $t-J$ model the d-wave BCS
294: wave function with a spin-rotationally invariant density-density 
295: Jastrow factor represents
296: a very accurate variational state, for the Hubbard model at small and 
297: intermediate coupling ($U \le 10t$) the best choice for the variational and
298: guiding wave function is given by a Jastrow-Slater determinant with rotated
299: orbitals Eqs.~(\ref{rotated}) and (\ref{rotated2}). Indeed, although the d-wave
300: BCS wave function is a singlet and does not break the SU(2) symmetry, it has
301: a very poor variational energy for the Hubbard model. 
302: The quality of the variational energy obtained with our 
303: Jastrow-Slater determinant remains considerably better than the 
304: BCS one at half-filling and $U \le 10t$ even when the accuracy 
305: of the approximation is improved by the GFMC.
306: Instead, in the doped case, the BCS wave function with GFMC 
307: is only slightly worse than the corresponding Jastrow-Slater determinant 
308: proposed in this work. This may suggest that antiferromagnetism is already 
309: suppressed at small finite doping, and d-wave superconductivity is a 
310: possible stable phase especially at large $U/t$. 
311: 
312: An important systematic improvement of the wave function can be achieved
313: by performing exactly one Lanczos step starting from $|\Psi_G \rangle$,
314: \begin{equation}
315: \label{LS}
316: |\Psi_L \rangle = \left ( 1+\alpha H \right ) |\Psi_G \rangle,
317: \end{equation}
318: with $\alpha$ free parameter chosen to minimize the energy.
319: This technique has been successfully used for the $t-J$ model both to improve
320: the variational calculation \cite{heeb} and as a starting point for power
321: methods \cite{chen}. Henceforth, we will denote by VMC and LS the results 
322: obtained with the wave function Eqs.~(\ref{afwf}) and (\ref{LS}), respectively, 
323: using variational Monte Carlo. Analogously, the symbols FN and FNLS will
324: indicate the fixed-node approximation applied to the wave function 
325: Eqs.~(\ref{afwf}) and (\ref{LS}), respectively.
326: Finally, the symbols SR will denote the stochastic reconfiguration approximation
327: applied to the wave function Eq.~(\ref{LS}).
328: 
329: In Tables \ref{tabella} and \ref{tabella2} 
330: we report the energies of 18 and 10 electrons on 18 sites, 
331: respectively. At half-filling we compare the results using Eq.~(\ref{afwf})
332: with the Slater determinant $|{\cal D} \rangle$ given by Eqs.~(\ref{det1}) 
333: and (\ref{det2}) for different
334: approximations and values of $U/t$. Using Eq.~(\ref{det2}) we obtain 
335: a sizeable improvement for large $U$'s ($U \ge 10t$). Notice that for $U=20t$
336: the best variational result (the FNLS) with Eq.~(\ref{det1}) is worse than the 
337: simple VMC with Eq.~(\ref{det2}). For 10 electrons the two Slater determinants
338: give the same results. Indeed, for this doping the antiferromagnetic order
339: is strongly suppressed and the Jastrow factor ${\cal J}$ does not play any
340: important role.
341: 
342: \section{Results}\label{results}
343: 
344: One of the most debated issues in strongly correlated electron models
345: is the nature of the charge distribution in their GS.
346: Recently, many authors \cite{lee,lin,putikka,hellberg,kohno,cbs,rommer} 
347: have addressed the question of PS in the $t-J$ model.
348: It is well accepted that for $J \gg t$, holes tend to cluster together
349: leaving the rest of the system in an antiferromagnetic state 
350: without holes. Although most of the calculations lead to the
351: conclusion that a critical value of $J$ below which
352: the GS is homogeneous exists, it is not clear what this value is at low
353: doping, ranging
354: from $0.5t$ and $1.2t$. Moreover, some authors \cite{cdw} have suggested
355: that just before PS, the GS has charge modulations.
356: Recently, an accurate numerical study of a few chains on the Hubbard
357: model \cite{bonca} has shown evidence of stripes, i.e., CDW oscillations.
358: However, this result appears limited to quasi-one-dimensional
359: geometry, as also suggested by the authors.
360: 
361: It is well known that PS is characterized by an infinite compressibility
362: in the thermodynamic limit. 
363: The compressibility can be related to the curvature of the energy
364: with respect to the electron density,
365: \begin{equation}\label{energy}
366: \chi = \left ( {\partial^2 E \over \partial n^2} \right )^{-1}.
367: \end{equation}
368: From the above definition we have 
369: that a divergent $\chi$ corresponds to a vanishing curvature of the 
370: energy as a function of density.
371: Therefore, PS can in principle be detected by means of energy measurements
372: for various densities. This is an appealing property,
373: since the accuracy on the energy is usually better than that of any other 
374: observable for most numerical methods.
375: Many previous numerical studies of PS have therefore concentrated
376: on the energy curve, or equivalently, on the energy per hole
377: $e_h(\delta)= [e(\delta) - e_H]/\delta$, 
378: where $e(\delta)$ is the energy per site at a hole density $\delta=1-n$ and
379: $e_H=e(0)$ is the energy at half-filling \cite{lin}. 
380: If $\chi$ diverges, $e_h(\delta)$ is flat in the thermodynamic limit
381: and develops a minimum for $\delta=\delta_c$ in finite systems, due
382: to the finite positive surface energy at the phase boundary.
383: 
384: This approach has been pursued by Cosentini {\it et al.} \cite{cosentini},
385: using FN calculations. They found that there is a large region of PS 
386: in the phase diagram, at least for
387: $U \ge 10t$, clearly in contrast with what is found in the $t-J$ model.
388: In this paper, we consider the Hubbard model and we show that a study
389: of PS instability is very difficult using only energy calculation.
390: Instead, a careful calculation of charge-correlation functions 
391: strongly indicates that the GS is homogeneous.
392: 
393: In order to show that the energy calculations may overestimate the
394: tendency to a PS instability, it is important to compare the GFMC 
395: results with some exact reference results.
396: Previous studies have shown 
397: that it is important to consider relatively large lattice
398: sizes since finite-size effects favor PS \cite{imada}.  
399: We need, therefore, a reference result for large lattices, 
400: where exact diagonalization is not available. 
401: In the case of the Hubbard model for small $U$, the AFQMC is almost 
402: exact and represents the reference we need.
403: As stated in the preceding section, we consider only closed shell doping 
404: because, at least for small $U$, huge finite size 
405: effects affect the physical properties in a rather drastic way.
406: For instance, the large bare density of states near half-filling 
407: determines an unphysical and spurious PS up to the first closed
408: shell \cite{imada}.
409: 
410: In Fig.~\ref{accuracy}, we show the accuracy of the GFMC results obtained 
411: with different approximations
412: compared with the AFQMC ones for a 162-site lattice
413: and $U=4t$. 
414: For this coupling value, AFQMC does not provide evidence for PS.
415: We plot $[E(\delta)-E_{hs}(\delta)]/E_{hs}(\delta)$,
416: where $E(\delta)$ and $E_{hs}(\delta)$ are the energies of GFMC and AFQMC,
417: respectively, for a doping $\delta$.
418: Besides the improvement in the absolute accuracy, 
419: the curves get flatter and flatter improving the
420: approximation, but only the SR accuracy is almost doping-independent. 
421: In other words, we need a very accurate calculation to 
422: eliminate the spurious dependence of the 
423: variational energy upon doping \cite{note}.
424: Even for the best variational method, the FNLS, although the 
425: accuracy on the energy is for all doping less than $1\%$, 
426: the difference in accuracy between, for example, the half-filled case
427: and the first closed shell is still sizable.
428: This difference is very important, because it represents just the energy scale
429: determining or ruling out PS.
430: 
431: In Fig.~\ref{emery}, the function $e_h(\delta)$ is shown for the
432: FN, FNLS, SR, and AFQMC methods.
433: We need to use SR to exclude the occurrence of PS, 
434: where even the FNLS data would imply PS. 
435: The reason for this disappointing situation is that all 
436: the known variational methods are still
437: too dependent on the guiding wave function. 
438: With the previous analysis, the resolution in energy necessary to detect 
439: or rule out PS is very hard
440: to reach with statistical methods, especially for large $U/t$.
441: 
442: On the other hand, GFMC methods have proven to be reliable
443: not only for energy calculations but also for correlation functions such as 
444: $N(q) = \langle n_q n_{-q}\rangle$,
445: where $n_q$ is the Fourier transform of the electron density \cite{cbs}.
446: For a phase-separated system, there are strong fluctuations in the density for
447: small momenta and $N(q \rightarrow 0)$ is expected to be strongly
448: enhanced for small momenta, that is, for $|q| \sim \frac{2\pi}{\xi}$, where
449: $\xi$ is the characteristic length of the phase-separated region.
450: Moreover, if $\chi$ diverges, also $N(q \rightarrow 0)$ diverges, yielding
451: an alternative tool to probe PS.
452: 
453: This method turns out to be more reliable, since it is 
454: based on a single calculation for a given doping value, whereas the
455: evaluation using $e_h(\delta)$ involves a comparison between energies obtained
456: by different simulations for different fillings, with the corresponding
457: guiding wave function having different accuracies.
458: Indeed with GFMC, $N(q)$ has been
459: proved to be a very sensitive tool to look at for detecting PS. 
460: In the $t-J$ model, $N(q)$ has a very different shape 
461: for stable and unstable systems. 
462: A clear peak at the smallest $q$ indicates PS even when, as shown in the
463: inset of Fig.~\ref{nqu4}, $J/t$ is very close to the PS boundary \cite{cbs}. 
464: Moreover, from $N(q)$ it is also possible to extract information about
465: charge fluctuations at finite $q$'s, related to CDW. 
466: In Ref.~\cite{cbs} we have
467: shown that in the $t-J$ model for $J=0.4t$, $N(q)$ has some peaks at
468: finite $q$'s, surprisingly near to what was found in recent experiments 
469: \cite{peak}. The knowledge of $N(q)$ allows us to extract more
470: general results with respect to the simple study of $e_h(\delta)$.
471: Furthermore, $N(q)$ is found to be much less size dependent than $e_h(\delta)$.
472: For the $t-J$ model there is no appreciable difference between 98-site
473: and 162-site lattices and for $J=0.4t$. 
474: Both calculations suggest that there is no
475: PS, whereas an analysis using the FN approximation of $e_h(\delta)$ should
476: lead to PS for 98 sites and to a homogeneous state for 162 sites.
477: Indeed, in this case PS is only a size effect and a homogeneous state
478: is found by increasing the accuracy of the method or 
479: by increasing the lattice size.
480: 
481: We computed $N(q)$ by means of the forward-walking technique \cite{calandra},
482: within the FNLS approximation, at half-filling
483: and for the first few closed-shell configurations on a 162- and a 
484: 98-site lattice. 
485: The evaluation of the density-density correlation function is in principle possible
486: even within SR by numerical differentiation of the energy with respect to an
487: external field coupled to $N(q)$. However, this approach is very
488: demanding and does not give a significant improvement on the FNLS 
489: results, which are very accurate. Indeed for the smallest $q$ vector 
490: for 90 electrons on the
491: 98-site lattice, i.e., $q=(2\pi /7,2\pi /7)$, we found the the AFQMC gives
492: $N(q)=0.0932(2)$ and the FNLS gives $N(q)=0.096(1)$. 
493:  
494: In Fig.~\ref{nqu4}, $N(q)$ is shown for $U=4t$ at half-filling for
495: a 162-site and a 98-site lattice
496: and for 154 electrons on a 162-site lattice and 90 electrons on 
497: a 98-site lattice.
498: No sign of divergence, and consequently of PS or CDW, is seen in the data.
499: We also notice that the two sets of points for the half-filled systems 
500: lie on the same curve, showing that we have 
501: substantially reached the thermodynamic limit.
502: For this value of $U$, $N(q)$ is essentially featureless for all 
503: doping we considered, suggesting that there are no charge
504: instabilities at any finite length. The smallest doping we considered
505: is $\delta \simeq 0.049$ and we cannot exclude that for smaller
506: doping PS or CDW are present. 
507: 
508: In order to investigate smaller doping,
509: we should consider larger lattices. Unfortunately, the accuracy of the
510: approximations considered decreases when increasing the size of the system
511: and the 162-site lattice represents
512: the largest lattice where the accuracy is acceptable. 
513: In Fig.~\ref{halffilling},
514: we report $[E(0)-E_{hs}(0)]/E_{hs}(0)$ for various sizes and for different
515: approximations: from the 18 sites to the 162 sites, the accuracy 
516: of FNLS changes from less than $0.1\%$ to about $0.5\%$. These indications
517: prevent us from considering sizes larger than the ones presented in this
518: paper.
519: 
520: Now we turn to larger Coulomb interactions and consider $U=10t$, 
521: where the AFQMC results are not reliable due to large
522: fluctuations.
523: In principle, GFMC techniques do not suffer from intrinsic limitations in the
524: large-coupling regime and it is possible to consider any value of $U$.
525: In practice we need an accurate knowledge of the nodes, i.e.,
526: an accurate guiding wave function.
527: Our choice, Eq.~(\ref{afwf}), with orbitals given by Eqs.~(\ref{rotated}) and
528: (\ref{rotated2}) is a very good approximation for the half-filling
529: case. In Table \ref{tabella}, we report the energies for various methods
530: for 18 electrons on 18 sites at different $U$'s. Although all the
531: approximations are quite size-dependent, the wave function becomes more
532: and more accurate by increasing the Coulomb potential.
533: Therefore, we expect that it also gives a good starting point at least
534: close to half-filling.
535: 
536: We present results for $U=10t$, for which previous FN calculations based on
537: $e_h(\delta)$ and a less accurate wave function \cite{cosentini} have
538: shown PS up to $\delta\simeq 0.15$. Indeed, if we use 
539: $e_h(\delta)$ as a probe
540: for PS, we find that the phase diagram shows a large instability region,
541: confirming the results of Ref.~\cite{cosentini}.
542: As for the $U=4t$ case, this instability is very likely to be a spurious
543: effect, a consequence of the different energy accuracy for different doping. 
544: This possibility, which cannot be proved without knowing the exact
545: energies at strong coupling (at present impossible), is instead
546: very clearly supported by the calculation of the charge-correlation
547: functions.
548: 
549: Figure \ref{nqu10} displays $N(q)$ for the same fillings of Fig.~\ref{nqu4}
550: and for 138 electrons on the 162-site lattice,
551: which corresponds to $\delta\simeq 0.148$.
552: All the correlation functions are definitely nondivergent for 
553: $q \rightarrow 0$ and are qualitatively similar to the $U=4t$ case,
554: indicating that the system is far away from a PS instability.
555: Furthermore $N(q)$ does not show peaks at any finite momenta 
556: for this Coulomb interaction.
557: This finding shows that the Hubbard and the $t-J$ model may have 
558: different behaviors as far as charge correlations are concerned.
559: Indeed, by diagonalizing exactly the 18-site lattice, we find that charge
560: fluctuations for the Hubbard and for the $t-J$ model are quite different
561: in the small doping region for $U=10t$ and $J=0.4t$, respectively \cite{becca}.
562: 
563: \section{Conclusions}\label{conclusions}
564: 
565: An extensive GFMC analysis of the Hubbard model at low hole doping
566: has been carried out. In particular, we have focused on the possible
567: instability of the model with respect to PS and CDW.
568: Comparing GFMC results with AFQMC ones in the weak-coupling region 
569: ($U=4t$), we show that detecting PS by means of
570: energy results requires a very accurate calculation at all electronic
571: densities.
572: Indeed, the accuracy of the energy is strongly
573: dependent on the electron density, and 
574: the signature of PS based only on energy calculations 
575: is clearly affected by this bias, 
576: leading to a spurious region of PS instability.
577: In the case of the Hubbard model, this is particularly relevant
578: because, while we are able to give a very good description of the
579: half-filled case, in which the GS is an antiferromagnetic insulator,
580: we are not aware of equally accurate descriptions of the doped state.
581: Even for $U=4t$, it is necessary to use the really accurate SR 
582: and AFQMC technique
583: to eliminate the doping dependence of the accuracy and to rule out PS.
584: 
585: On the other hand, PS (and CDW) instability can be probed more easily
586: using charge correlation functions. This approach has various advantages.
587: First, it is found that $N(q)$ has very small size effects and the
588: thermodynamic limit is reached with about 100 sites, both for the Hubbard
589: and the $t-J$ model. Second, the information contained in 
590: $N(q)$ does not depend on different densities, implying that a different
591: accuracy as a function of doping does not introduce any external bias.
592: 
593: Instead of using energy calculations, which are very expensive at moderate
594: and large $U$'s, we calculate the charge correlation functions and we
595: are able to find clear evidence for the absence of PS up to
596: $U=10t$ in the low doping regime.
597: 
598: \acknowledgments
599: 
600: It is a pleasure to acknowledge useful discussions with L. Capriotti, 
601: M. Calandra, G. Bachelet, E. Koch, and A. Parola, to whom we are also 
602: grateful for critical and careful reading of the manuscript.
603: This work has been partially supported by MURST-COFIN99 and by
604: Istituto Nazionale per la Fisica della Materia.
605: 
606: \begin{figure}
607: \centerline{\psfig{bbllx=50pt,bblly=250pt,bburx=500pt,bbury=650pt,%
608: figure=fig1.ps,width=80mm,angle=0}}
609: \caption{\baselineskip .185in \label{accuracy}
610: Relative accuracy of various GFMC techniques with respect to AFQMC
611: for a 162-site lattice with $U=4t$ as a function of filling $\delta$.
612: From top to bottom VMC (empty triangles), 
613: FN (empty hexagons), LS (empty squares),
614: FNLS (empty circles), and SR (full squares). Lines are guides to the eye.}
615: \end{figure}
616: 
617: \begin{figure}
618: \centerline{\psfig{bbllx=50pt,bblly=250pt,bburx=500pt,bbury=650pt,%
619: figure=fig2.ps,width=80mm,angle=0}}
620: \caption{\baselineskip .185in \label{emery}
621: Energy per hole $e_h(\delta)$ for a 162-site lattice with $U=4t$.
622: From top to bottom FN (empty hexagons), FNLS (empty circles), 
623: SR (full squares) , and AFQMC (full circles). Lines are guides to the eye.}
624: \end{figure}
625: 
626: \begin{figure}
627: \centerline{\psfig{bbllx=50pt,bblly=250pt,bburx=500pt,bbury=650pt,%
628: figure=fig3.ps,width=80mm,angle=0}}
629: \caption{\baselineskip .185in \label{nqu4}
630: $N(q)$ for $U=4t$, 162 electrons on 162 sites (empty circles), 98 electron
631: on 98 sites (empty squares), 154 electrons on 162 sites (full circles)
632: and 90 electrons on 98 sites (full squares). Lines are guides to the eye and
633: error bars are smaller than points.  $\Gamma = (0,0)$, $X = (\pi,\pi)$, $M = (\pi,0)$.
634: In the inset: $N(q)$ for the $t-J$ model, $J=0.6t$, 156 electrons on 162 sites 
635: (squares), and 94 electrons on 98 sites (circles).}
636: \end{figure}
637: 
638: \begin{figure}
639: \centerline{\psfig{bbllx=50pt,bblly=250pt,bburx=500pt,bbury=650pt,%
640: figure=fig4.ps,width=80mm,angle=0}}
641: \caption{\baselineskip .185in \label{halffilling}
642: Relative accuracy of various GFMC techniques with respect to AFQMC
643: for different lattices ($L=18,98,162$) and $U=4t$.
644: From top to bottom: VMC (empty triangles), 
645: FN (empty hexagons), LS (empty squares),
646: and FNLS (empty circles). Lines are guides to the eye.}
647: \end{figure}
648: 
649: \begin{figure}
650: \centerline{\psfig{bbllx=50pt,bblly=250pt,bburx=500pt,bbury=650pt,%
651: figure=fig5.ps,width=80mm,angle=0}}
652: \caption{\baselineskip .185in \label{nqu10}
653: $N(q)$ for $U=10t$, 162 electrons on 162 sites (empty circles), 98 electrons
654: on 98 sites (empty squares), 154 electrons on 162 sites (full circles),
655: 90 electrons on 98 sites (full squares), and 138 electrons on 162 sites 
656: (full triangles). Lines are guides to the eye and error bars are smaller
657: than points.
658: $\Gamma = (0,0)$, $X = (\pi,\pi)$, $M = (\pi,0)$.}
659: \end{figure}
660: 
661: \begin{table}
662: \begin{tabular}{lllllll}
663: $U$   & $|{\cal D} \rangle $ & $E_{ex}$ & $E_{VMC}$  & $E_{FN}$   & $E_{LS}$   & $E_{FNLS}$  \\
664: \hline \hline
665: 4t    &  (\ref{det1})        & -0.9585 & -0.9382(1) & -0.9514(1) & -0.9520(1) & -0.9556(1) \\
666: 10t   &  (\ref{det1})        & -0.4484 & -0.4034(1) & -0.4284(1) & -0.4154(1) & -0.4316(1) \\
667: 20t   &  (\ref{det1})        & -0.2339 & -0.2023(1) & -0.2195(1) & -0.2060(1) & -0.2225(1) \\
668: 4t    &  (\ref{det2})        & -0.9585 & -0.9460(1) & -0.9547(1) & -0.9553(1) & -0.9576(1) \\
669: 10t   &  (\ref{det2})        & -0.4484 & -0.4382(1) & -0.4451(1) & -0.4428(1) & -0.4470(1) \\
670: 20t   &  (\ref{det2})        & -0.2339 & -0.2293(1) & -0.2232(1) & -0.2310(1) & -0.2337(1) 
671: \end{tabular}
672: \caption{GS energies for 18 electrons on 18 sites as a function of $U/t$ using (\ref{det1})
673: and (\ref{det2}) as Slater determinant.}
674: \label{tabella}
675: \end{table}
676: 
677: \begin{table}
678: \begin{tabular}{llllll}
679: $U$   & $E_{ex}$ & $E_{VMC}$  & $E_{FN}$   & $E_{LS}$   & $E_{FNLS}$  \\
680: \hline \hline
681: 4t    & -1.1299 & -1.1124(1) & -1.1218(1) & -1.1229(1) & -1.1263(1) \\
682: 10t   & -1.0193 & -0.9749(1) & -1.0006(1) & -0.9997(1) & -1.0098(1) \\
683: 20t   & -0.9598 & -0.8983(1) & -0.9354(1) & -0.9253(1) & -0.9450(1)
684: \end{tabular}
685: \caption{GS energies for 10 electrons on 18 sites as a function of $U/t$.}
686: \label{tabella2}
687: \end{table}
688: 
689: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
690: %                           BIBLIOGRAPHY
691: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
692: 
693: \begin{thebibliography}{99}
694: 
695: \bibitem{hubbard} J. Hubbard, Proc. R. Soc. London, Ser. A {\bf 276}, 
696:    238 (1963); M.C. Gutzwiller, \prl {\bf 10}, 159 (1963); 
697:    J. Kanamori, Prog. Theor. Phys. {\bf 30}, 275 (1963).
698: \bibitem{jorgensen} J.D. Jorgensen, B. Dabrowski, Shiyon Pei, D.G. Hinks,
699:    L. Soderholm, B. Morosin, J.E. Shirber, E.L. Venturini, and 
700:    D.S. Ginley, \prb {\bf 38}, 11337 (1988).
701: \bibitem{tranquada} J.M. Tranquada, B.J. Sternlib, J.D. Axe, Y. Nakamura,
702:    and S. Uchida, Nature {\bf 375}, 561 (1995).
703: \bibitem{emery} V.J. Emery and S.A. Kivelson, Physica C {\bf 209}, 
704:    597 (1993).
705: \bibitem{castellani} C. Castellani, C. Di Castro, and M. Grilli, \prl {\bf 75},
706:    4650 (1995).
707: \bibitem{white} S.R. White, \prl {\bf 69}, 2863 (1992); S.R. White, \prb
708:    {\bf 48}, 10345 (1993).
709: \bibitem{afqmc} D.R. Hamann and S.B. Fahy, \prb {\bf 41}, 11352 (1990);
710:    S.B. Fahy and D.R. Hamann, \prl {\bf 65}, 3437 (1990).
711: \bibitem{nandini} N. Trivedi and D.M. Ceperley, \prb {\bf 41}, 4552 (1990).
712: \bibitem{cpqmc} Shiwei Zhang, J. Carlson, and J.E. Gubernatis, \prl {\bf 74},
713:    3652 (1995); Shiwei Zhang, J. Carlson, and J.E. Gubernatis, \prb {\bf 55},
714:    7464 (1997).
715: \bibitem{fn} D.F.B. ten Haaf, H.J.M. van Bemmel, J.M.J. van Leeuwen,
716:    W. van Saarloos, and D.M. Ceperley, \prb {\bf 51}, 13039 (1995).
717: \bibitem{sorella} S. Sorella, \prl {\bf 80}, 4558 (1998).
718: \bibitem{sorcap} S. Sorella and L. Capriotti, \prb {\bf 61}, 2599 (2000).
719: \bibitem{franjic} F. Franjic and S. Sorella, Prog. Theor. Phys. {\bf 97},
720:    399 (1997).
721: \bibitem{shiba} H. Yokoyama and H. Shiba, J. of Phys. Soc. of Japan. 
722:    {\bf 56}, 3582 (1987).
723: \bibitem{cosentini} A. Cosentini, M. Capone, L. Guidoni, and G. Bachelet,
724:    \prb {\bf 58}, 14685 (1998).
725: \bibitem{calandra} M. Calandra and S. Sorella, \prb {\bf 57}, 11446 (1998).
726: \bibitem{heeb} E.S. Heeb and T.M. Rice, Europhys. Lett. {\bf 27}, 673 (1994).
727: \bibitem{chen} Y.C. Chen and T.K. Lee, \prb {\bf 51}, 6723 (1995).
728: \bibitem{lee} C.T. Shih, Y.C. Chen, and T.K. Lee, \prb {\bf 57}, 627 (1998).
729: \bibitem{lin} V.J. Emery, S.A. Kivelson, and H.Q. Lin, \prl {\bf 64},
730:    475 (1990).
731: \bibitem{putikka} W.O. Putikka, M.U. Lucchini, and T.M. Rice, \prl {\bf 68},
732:    538 (1992).
733: \bibitem{hellberg} C.S. Hellberg and E. Manousakis, \prl {\bf 78}, 
734:    4609 (1997).
735: \bibitem{kohno} M. Kohno, \prb {\bf 55}, 1435 (1997).
736: \bibitem{cbs} M. Calandra, F. Becca, and S. Sorella, \prl {\bf 81}, 5185 
737:    (1998).
738: \bibitem{rommer} S. Rommer, S.R. White, and D.J. Scalapino, \prb {\bf 61}, 13424 (2000).
739: \bibitem{cdw} S.R. White and D.J. Scalapino, \prl {\bf 80}, 1272 (1998);
740:    {\bf 81}, 3227 (1998).
741: \bibitem{bonca} J. Bonca, J.E. Gubernatis, M. Guerrero, 
742:    E. Jeckelmann, and S.R. White, \prb {\bf 61}, 3251 (2000).
743: \bibitem{imada} N. Furukawa and M. Imada, J. of Phys. Soc. of Japan
744:    {\bf 61}, 3331 (1992).
745: \bibitem{note} We checked that without rotating orbitals, that is,
746:    using the wave function of Ref.~\cite{cosentini}, 
747:    we have much less accuracy as a function of density.
748: \bibitem{peak} J.M. Tranquada, J.D. Axe, N. Ichikawa, A.R. Moodenbaugh,
749:    Y. Nakamura, and S. Ucida, \prl {\bf 78}, 338 (1997).
750: \bibitem{becca} F. Becca, A. Parola, and S. Sorella, \prb {\bf 61}, 16287 (2000).
751: \end{thebibliography}
752: 
753: \end{document}
754: